You are on page 1of 10

Tribol Lett (2014) 56:133–142

DOI 10.1007/s11249-014-0392-2

ORIGINAL PAPER

Effect of Graphene and Ionic Liquid Additives on the Tribological


Performance of Epoxy Resin
N. Saurı́n • J. Sanes • M. D. Bermúdez

Received: 4 June 2014 / Accepted: 6 August 2014 / Published online: 21 August 2014
Ó Springer Science+Business Media New York 2014

Abstract In the present study, we have determined the Keywords Epoxy resin  Ionic liquid  Graphene 
effect of graphene (PG), the ionic liquid (IL) and PG Nanocomposites  Dynamic friction  Wear and lubrication
modified by mechanical blend with the IL 1-octyl-3-
methylimidazolium tetrafluorobotate (IL ? PG) on the
tribological performance of epoxy resin (ER). IL ? PG 1 Introduction
stable suspensions have been added to an epoxy resin (ER)
matrix to obtain the new nanocomposite (ER ? IL ? PG), Tribology of polymers and polymer-matrix composites is
and its tribological performance has been compared with increasingly relevant due to their growing applications in
that of neat epoxy resin and with the nanocomposites sliding parts, as surface finishing materials and in self-
containing PG (ER ? PG) or IL (ER ? IL). While neat ER lubricating components [1–5]. Polymer composites con-
presents a high dynamic friction coefficient of up to 0.31 taining graphene (PG) are an active research field due to their
and a severe wear with a specific wear rate of 8.1 9 10-4 improved mechanical, thermal and electrical properties [6–
mm3 N-1m-1, the new nanocomposites show negligible 11]. Epoxy resin nanocomposites containing PG or graphene
surface damage, as determined by surface roughness and oxide are currently the object of great attention due to their
profilometry. All nanocomposites show low friction coef- enhanced mechanical or electrical performance [12].
ficients and negligible wear. The maximum friction Room temperature ionic liquids (ILs) have shown an
reduction, up to a 70 %, is obtained for ER ? PG. Results outstanding lubricating performance for metals, polymers
are discussed upon the basis of TEM microscopy, SEM and ceramic materials under severe conditions [13–18] due
microscopy and EDX analysis, differential scanning calo- to their ability to form ordered adsorbed tribolayers.
rimetry, thermogravimetric analysis and dynamic We have previously described [19, 20] the friction and
mechanical analysis. Addition of IL or IL ? PG has a wear reduction of epoxy resin by external lubrication or by
plasticizing effect on ER, while addition of PG increases addition of ILs. A very recent research [21] has shown that
the thermal stability and stiffness of ER. PG shifts the ILs can act as cross-linkers of epoxy resin and as disper-
storage modulus onset and in the loss modulus and tan d sants of graphite nanoplatelets in an epoxy matrix.
maximum peaks to higher temperatures, while a shift to Graphene, constituted by one-atom thick layers of sp2
lower values is observed with addition of IL or IL ? PG. carbon, has risen and extraordinary scientific and technical
interest [22, 23]. The tribological and surface science
applications of PG are some of the most front edge lines of
research at the present moment.
There exist very few precedents of the use of PG in
N. Saurı́n  J. Sanes  M. D. Bermúdez (&) friction and wear reduction [23, 24].
Grupo de Ciencia de Materiales e Ingenierı́a Metalúrgica, Choudhary et al. [25] have reported a 26 % friction
Departamento de Ingenierı́a de Materiales y Fabricación,
reduction and a 9 % wear reduction for steel by lubrication
Universidad Politécnica de Cartagena, Campus de la Muralla del
Mar. C/Doctor Fleming, s/n, 30202 Cartagena, Spain with hexadecane containing chemically modified PG, which
e-mail: mdolores.bermudez@upct.es is supposed to intercalate between the sliding surfaces.

123
134 Tribol Lett (2014) 56:133–142

Fig. 1 TEM micrographs: a,


b as-received graphene (PG);
c IL ? PG; d ER ? PG;
e ER ? IL ? PG

Kandanur et al. [26] have shown that the wear rate of the PG sheets could reduce wear by inhibiting crack
polytetrafluoroethylene (PTFE) is reduced in four orders of propagation.
magnitude by addition of a 10 % of PG sheets of 3–4 layers The present work was carried out in the expectation that
as nanoreinforcements. These authors have proposed that the reinforcing effect of PG and the dispersing ability of

123
Tribol Lett (2014) 56:133–142 135

PG was added to the prepolymer) were processed as previ-


1.0 IL + PG ously described [19, 20]. In order to obtain ER ? 9 wt%
PG
IL ? 0.1 % PG (ER ? IL ? PG), PG (0.1 g) was added to
0.8 IL (9.0 g) and the mixture was mechanically stirred in an
agate mortar for 5 min and then at 1,600 rpm for 30 s to
Normalized intensity

0.6
obtain IL ? PG which was added to the prepolymer before
processing following the method described [19, 20].
SEM images were obtained using a Hitachi S3500N.
0.4
The epoxy samples were platinum-coated with a SC7640
Sputter Coater before observation. TEM micrographs were
0.2 obtained with a high-resolution JEOL JEM 2100. Raman
spectra were collected using a Nicolet Almega XR Raman
0.0
spectrometer from Thermo Electron, equipped with an
0 1000 2000 3000 Olympus microscope. A laser of 514 nm at 100 % of
-1
Wavenumber (cm ) power and 50 lm of aperture was used. A differential
scanning calorimeter (DSC) 822e (Mettler-Toledo) was
Fig. 2 Raman spectra of as-received PG and IL-modified graphene used to determine glass transition temperatures (Tg); 15 mg
(IL ? PG)
samples were heated under an inert N2 atmosphere
(50 ml min-1), at a heating rate of 10 °C min-1. Degra-
Table 1 Raman spectroscopy data dation temperatures (Td) were determined in air with a
Material D band G band 2D band ID/IG I2D/IG Shimadzu thermogravimetric analysis (TGA)-50 equip-
(cm-1) (cm-1) (cm-1) ment (20 °C min-1; N2, 40 ml min-1). Dynamic mechan-
PG 1,358.3 1,581.8 2,721.8 0.16 0.52
ical analysis (DMA) under the dual-cantilever clamping
mode was carried out on prismatic test specimens
IL ? PG 1,356.7 1,583.3 2,728.3 0.28 0.52
(dimensions 17 9 3.8 9 3.3 mm) with a TA Q800 DMA
analyser equipped with a TA Universal Analysis 2000
ILs could also improve the poor tribological performance software. Glass transition temperature (Tg) values were
of epoxy resin. determined at an oscillatory frequency of 1 Hz, under a
Chemical modification of PG has been described by 1 % strain, for a temperature range from 30 to 120 °C at a
covalent and non-covalent interactions. However, covalent heating rate of 3 °C min-1.
modifications can modify the PG conjugation system and, Pin-on-disc tests were carried out with a ISC 200-PC
consequently, its properties. Non-covalent modification by pin-on-disc tribometer using discs (diameter 40 mm;
p–p stacking and van der Waals interactions is believed to thickness 4 mm) against AISI 316L stainless steel balls
maintain the structure and properties of PG [27]. In the (hardness 210 HV) with a 0.8 mm sphere radius, under dry
present study, non-covalent interactions between PG and conditions, that is, without external lubrication. Normal
IL have been produced under mild conditions. applied load was 0.49 N (mean contact pressure 0.07 GPa),
The IL selected for the present study combines a high sliding velocity 0.10 ms-1, sliding distance 500 m and
thermal stability with a high molecular polarity, which allows sliding radius 9 mm. The 500 m sliding distance allows a
the formation of adsorbed layers on metal surfaces, and the steady state regime to be reached for all materials studied.
presence of a long alkyl side chain which could improve the Wear rates were determined by means of a Talysurf CLI
compatibility of nanophases with the polymer network. profilometer. At least three tests were performed under the
To the best of our knowledge, the present study is the same experimental conditions for each material, at room
first on the tribological properties of epoxy resin-PG or temperature with a relative humidity of 40 ± 5 %. Steel
epoxy resin IL-modified-PG nanocomposites. balls were observed by optical microscopy (Leica DMRX).

2 Experimental
3 Results and Discussion
1–2 Layers PG (Avanzare Nanotechnology; Spain) and the
IL 1-octyl-3-methylimidazolium tetrafluoroborate (Fluka, 3.1 TEM Microscopy
Germany) ([98 % purity) were used as-received. ER and
ER ? 9 wt% IL (ER ? IL) were prepared following a pre- Figure 1a shows a TEM micrograph of the as-received PG
viously described method [19]. ER ? PG (where a 0.1 wt% sheets. Figure 1b shows the ordered structure present in the

123
136 Tribol Lett (2014) 56:133–142

C 1s
285.0

286.7

686.1
401.95
F1s N1s

194.2 532.4
B1s O1s

Fig. 3 XPS analysis of neat IL

ionic liquid-modified graphene IL ? PG. The different 3.3 XPS Analysis


stacking of PG layers in Fig. 1d as compared to Fig. 1c
could be attributed to a possible exfoliating effect of the IL Figure 3 shows the binding energies for C1s, F1s, N1s and
on PG [28]. ER ? PG (Fig. 1e) shows the presence of PG B1s present in neat IL. The O1s peak observed at 532.4 eV
platelets similar to those present in PG (Fig. 1b). could be attributed to water impurities present in the IL.
Figure 4 shows the corresponding binding energies for
3.2 Raman Spectroscopy IL ? PG. F1s, N1s and B1s peaks are similar to those found
for neat IL and show the surface modification of PG by IL.
Figure 2 shows the G band, characteristic of in-plane The main differences are observed for C1s and O1s peaks.
vibrations of the graphitic wall, at 1,581.8 cm-1 (Table 1). Neat IL shows a C1s peak at 285.0 eV assignable to the
The D band for PG is observed at 1,358.3 cm-1 (Table 1). aliphatic carbon (and also to carbon impurities) atoms
The low ID/IG intensity ratio of 0.16 is indicative of a rela- present in the alkyl substituents of the imidazolium ring. The
tively low proportion of defects. The 2D band, characteristic same peak is present in IL ? PG. The C1s peak of neat IL at
of PG, which originates from second-order Raman scattering 286.7 eV is assignable to the sp2 carbon atoms of the imi-
process [29], appears at 2,721.8 cm-1. Graphene bands in dazolium ring. In the case of IL ? PG, two peaks assignable
the ER ? PG nanocomposite cannot be analysed due to the to sp2 carbon atoms are observed at 286.2 and 287.4 eV,
absorptions of the epoxy resin in the same region. respectively, which could be assigned to the imidazolium

123
Tribol Lett (2014) 56:133–142 137

Fig. 4 XPS analysis of


IL ? PG
C 1s 2 285.0

286.2
287.4

686.8 402.7
F1s N1s

194.85
533.9
B1s O1s

532.2

ring and PG, respectively. A new O1s peak is observed in 3.5 Thermal Properties
IL ? PG at 533.9 eV. This could be tentatively assigned to
graphene oxide residues present in the commercial PG used. Table 2 compares the thermal properties [30] of the new
nanocomposites with those of the epoxy resin. ER ? PG
presents a glass transition temperature of 90.3 °C, similar
3.4 Nanophase Distribution to that of ER (91.9 °C), and shows a higher thermal sta-
bility, with a degradation temperature (for a 50 % weight
Previous studies [19] on ER ? IL nanocomposites have loss) of 348.4, 4.5 °C higher than that of ER (343.9 °C). In
shown a uniform distribution of ILs in the epoxy matrix. contrast, the addition of IL lowers the glass transition
Figure 5 shows the fracture surface of ER ? IL (Fig. 5a) temperature both in ER ? IL and ER ? IL ? PG, while
and ER ? IL ? PG (Fig. 5b). In both cases, the fluorine maintaining a similar or even higher degradation temper-
element maps show a uniform distribution of the additives ature, due to the high thermal stability of the additives.
in small rounded particles. In the case of ER ? IL ? PG, These results show that IL or IL ? PG have a plasticizing
the uniform distribution is attributed to surface modifica- effect on the epoxy resin, thus increasing chain mobility.
tion and partial exfoliation of PG by IL, although some The experimental conditions used in the present case
larger size agglomerates are also present (Fig. 5b). were selected in order to compare the thermal behaviour of

123
138 Tribol Lett (2014) 56:133–142

Fig. 5 SEM micrograph and


fluorine element map of the
fracture surface of: a ER ? IL;
b ER ? IL ? PG

Table 2 Thermal properties those of the neat epoxy resin (Fig. 6). The addition of a
0.1 wt% PG to epoxy resin to give ER ? PG increases
Material Glass transition Degradation (50 % weight
temperature Tg (°C) loss) temperature, Td (°C)
the onset of the storage modulus from 940.6 MPa
(DSC) (TGA) (60.45 °C) to 1,126.0 MPa (64.96 °C) (Fig. 6a). In
contrast, the addition of IL (ER ? IL) reduces the
ER 91.9 343.9 storage modulus to 820.3 MPa (48.61 °C). Loss modu-
ER ? PG 90.3 348.5 lus (Fig. 6b) and tan d (Fig. 6c) peaks are both shifted
ER ? IL ? PG 68.7 327.6 to lower values by addition of IL. ER ? IL ? PG
ER ? IL 68.5 352.5 shows a storage modulus similar to that of ER due to
the opposite effect of the additives present in the
the new blends with that of the neat epoxy resin. Further nanocomposite. These results confirm the reinforcement
studies on thermal stability using low scanning rates or effect of PG even in a very low concentration, while
isothermal conditions [31] will be carried out in order to the presence of the fluid phase in ER ? IL results in a
determine their long-term stability. lower stiffness.
The higher stiffness of ER ? PG could anticipate a
3.6 Dynamic Mechanical Properties good tribological performance by preventing the crack
propagation and fracture mechanisms which are readily
The dynamic mechanical properties of the new nano- produced in neat epoxy resin, as will be discussed in the
composites have been determined and compared with following section.

123
Tribol Lett (2014) 56:133–142 139

(a) (a)
10000 0.18
ER
ER + PG 0.16
ER + IL
Storage modulus E' (MPa)

ER + IL + PG 0.14
1000

Friction coefficient
0.12

0.10
100
0.08

0.06
10
0.04

0.02
1
20 40 60 80 100 120 0.00
ER ER+PG ER+IL ER+IL+PG
Temperature (ºC)
(b)
(b) 0.35
350 ER
ER 0.30 ER + PG
300 ER + PG ER + IL + PG

Friction coefficient
ER + IL ER + IL
0.25
Loss modulus E'' (MPa)

ER + IL + PG
250
0.20
200
0.15
150
0.10
100

50 0.05

0 0.00
0 100 200 300 400 500
-50 Distance (m)
20 40 60 80 100 120
Temperature (ºC) Fig. 7 a Mean friction coefficient values after three tests; b repre-
sentative friction coefficient–distance curves
(c)
1.4
ER
ER + PG
1.2
ER + IL
materials containing IL show a very similar behaviour,
1.0
ER + IL + PG with a friction reduction of a 27 % with respect to ER.
Even more significant is the observation of the real-time
0.8
friction–distance curves (Fig. 7b). ER presents a maximum
tan Delta

0.6 friction value of 0.31 after 200 m, to reach a steady state


friction value of 0.27 until the end of the test. ER ? PG
0.4
maintains the lowest friction along the test. A maximum
0.2 friction reduction of a 70 % is obtained for ER ? PG with
0.0 respect to ER. The main advantage in the performance of
ER ? PG with respect to the rest of materials is the
-0.2
20 40 60 80 100 120 140 absence of the initial friction increasing period before
Temperature (ºC) reaching the stationary regime.
The similar behaviour of ER ? IL and ER ? IL ? PG
Fig. 6 Dynamic mechanical properties of ER and the nanocompos- could be attributed to the lubricating performance of IL,
ites: a storage modulus; b loss modulus; c tan d
which predominates over the effect of PG, due to its much
higher relative proportion.
ER shows a very severe surface damage after sliding
3.7 Tribological Performance against the steel ball (Fig. 8a), with a wide wear track
(wear rate 8.1 9 10-4 mm3 Nm-1) covered by wear
Figure 7a shows the mean friction coefficients for the debris.
epoxy resin and the nanocomposites. The maximum fric- As the nanocomposites present no wear rate, surface
tion reduction, of a 53 %, is obtained for ER ? PG. The roughness has been measured in order to quantify surface

123
140 Tribol Lett (2014) 56:133–142

Fig. 8 SEM micrographs and


surface topography profiles after
the tribological tests: a ER;
b ER ? PG; c ER ? IL;
d ER ? IL ? PG

changes with the tribological tests. SEM and profilometry high-load-carrying ability of the new nanocomposite.
observations (Fig. 8) confirm the wear prevention effect of ER ? IL (Fig. 8c) shows a maximum increase in surface
the additives with respect to neat ER (Fig. 8a). ER ? PG roughness from 0.06 to 0.26 lm, due to the presence of
(Fig. 8b) maintains a constant value of 0.14 lm on the microcracks which form in the absence of the reinforce-
contact region, before and after the test. This confirms the ment effect of PG. ER ? IL ? PG (Fig. 8d) shows a mild

123
Tribol Lett (2014) 56:133–142 141

polishing effect with a reduction of roughness from 2. Dasari, A., Yu, Z.Z., Mai, Y.W.: Fundamental aspects and recent
0.09 lm before the test to 0.05 lm after the test. progress on wear/scratch damage in polymer nanocomposites.
Mater. Sci. Eng. R 63, 31–80 (2009)
SEM and profilometry observation confirm the absence 3. Brostow, W., Deborde, J.L., Jaklewicz, M., Olszynski, P.: Tri-
of wear on ER ? PG, which is free from cracks and wear bology with emphasis on polymers: friction, scratch resistance
debris, as compared to ER (Fig. 8b). A similar absence of and wear. J. Mater. Educ. 25, 119–132 (2003)
surface damage is observed for ER ? IL (Fig. 8c) and 4. Myshkin, N.K., Petrokovets, M.I., Kovalev, A.V.: Tribology of
polymers: friction, wear and mass transfer. Tribol. Int. 38,
ER ? IL ? PG (Fig. 8d). 910–921 (2005)
5. Brostow, W., Kovacevic, V., Vrsaljko, D., Whitworth, J.: Tri-
bology of polymers and polymer based composites. J. Mater.
Educ. 32, 273–290 (2010)
4 Conclusions 6. Huang, X., Qi, X., Boey, F., Zhang, H.: Graphene-based com-
posites. Chem. Soc. Rev. 41, 666–686 (2012)
7. Potts, J.R., Dreyer, D.R., Bielawski, C.W., Ruoff, R.S.: Graph-
Addition of 0.1 wt% PG to epoxy resin shifts the onset of
ene-based polymer nanocomposites. Polymer 52, 5–25 (2011)
storage modulus, and the maximum of loss modulus and 8. Verdejo, R., Bernal, M.M., Romasanta, L.J., Lopez-Manchado,
tan d to higher temperature values, while reducing the M.A.: Graphene filled polymer nanocomposites. J. Mater. Chem.
friction coefficient of epoxy resin against stainless steel, 21, 3301–3310 (2010)
9. Kuilla, T., Bhadra, S., Yao, D.H., Kim, N.H., et al.: Recent
and preventing surface damage or wear.
advances in graphene based polymer composites. Prog. Polym.
1-Octyl-3-methylimidazolium tetrafluoroborate ionic Sci. 35, 1350–1375 (2010)
liquid shows an outstanding friction-reducing and antiwear 10. Kim, H., Abdala, A.A., Macosko, C.W.: Graphene/polymer
performance for the epoxy resin-stainless steel contact, nanocomposites. Macromolecules 43, 6515–6530 (2010)
11. Ramanathan, T., Abdala, A.A., Stankovich, S., Dikin, D.A., et al.:
which can be further improved by addition of a 0.1 wt%
Functionalized graphene sheets for polymer nanocomposites.
proportion of PG. Nat. Nanotechnol. 3, 327–331 (2008)
It is suggested that the reinforcement effect of PG would 12. Tang, L.C., Wan, Y.J., Yan, D., Pei, Y.B., et al.: The effect of
be responsible for the good tribological performance, while graphene dispersion on the mechanical properties of graphene/
epoxy composites. Carbon 60, 16–27 (2013)
the increase in polymer chain mobility and the ability of IL
13. Torimoto, T., Tsuda, T., Okazaki, K., Kuwawata, S.: New fron-
molecules to form ordered adsorbed tribolayers would tiers in materials science opened by ionic liquid. Adv. Mater. 22,
account for the good results obtained for ER ? IL. 1196–1221 (2010)
The nanocomposite-containing PG surface modified by 14. Minami, I.: Ionic liquids in tribology. Molecules 14, 2262–2269
(2009)
the ionic liquid shows a uniform distribution of the
15. Zhou, F., Liang, Y., Liu, W.: Ionic liquid lubricants: designed
nanophase, with similar thermal, dynamic mechanical and chemistry for engineering applications. Chem. Soc. Rev. 28,
tribological properties to that of epoxy resin ? ionic 2590–2599 (2009)
liquid material, where the ionic liquid acts as a plasti- 16. Bermúdez, M.D., Jiménez, A.E., Sanes, J., Carrión, F.J.: Ionic
liquids as advanced lubricants. Molecules 14, 2888–2908 (2009)
cizing agent, increasing chain mobility, reducing glass
17. Palacio, M., Bhushan, B.: A review of ionic liquids for green
transition temperature and shifting the storage modulus molecular lubrication in nanotechnology. Tribol. Lett. 40,
onset and the loss modulus and tan d peaks to lower 247–268 (2010)
temperatures. 18. Somers, A.E., Howlett, P.C., MacFarlane, D.R., Forsyth, M.: A
review of ionic liquid lubricants. Lubricants 1, 3–21 (2013)
The combination of PG and ionic liquid could be used in
19. Sanes, J., Carrión-Vilches, F.J., Bermúdez, M.D.: New epoxy-
the development of new nanolubricants and nanocompos- ionic liquid dispersions. Room temperature ionic liquid as
ites with tailored thermal, mechanical and tribological lubricant of epoxy resin-stainless steel contacts. E-Polym. (2007),
behaviour, following the simple method described herein. Art. No. 005
20. Sanes, J., Carrión, F.J., Bermúdez, M.D.: Effect of the addition of
Further studies are needed in order to determine the long-
room temperature ionic liquid and ZnO nanoparticles on the wear
term stability and tribological performance of the disper- and scratch resistance of epoxy resin. Wear 268, 1295–1302
sions and nanocomposites. (2010)
21. Maka, H., Spychaj, T., Kowalczyk, K.: Imidazolium and deep
Acknowledgments The authors thank the financial support of the eutectic ionic liquids as epoxy resin crosslinkers and graphite
Ministerio de Economı́a y Competitividad (MINECO, Spain) nanoplatelets dispersants. J. Appl. Polym. Sci. 131, 40401 (2014)
(MAT2011-23162). N. Saurin is grateful to MINECO (Spain) for a 22. Stankovich, S., Dikin, D.A., Dommett, G.H., Kohlhaas, K.M.,
FPI research Grant (BES-2012-056621). Zimney, E.J., Stach, E.A., Piner, R.D., Nguyen, S.T., Ruoff, R.S.:
Graphene-based composite materials. Nature 442, 282–286 (2006)
23. Berman, D., Erdemir, A., Sumant, A.V.: Graphene: a new
emerging lubricant. Mater. Today 17, 31–42 (2014)
References 24. Penkov, O., Kim, H.J., Kim, D.E.: Tribology of graphene: a
review. Int. J. Precis Eng. Manuf. 15, 577–585 (2014)
1. Burris, D.L., Boesl, B., Bourne, G.R., Sawyer, W.G.: Polymeric 25. Choudhary, S., Mungse, H.P., Khatri, O.P.: Dispersion of alkyl-
materials for tribological applications. Macromol. Mater. Eng. ated graphene in organic solvents and its potential for lubrication
292, 387–402 (2007) applications. J. Mater. Chem. 22, 21032–21039 (2012)

123
142 Tribol Lett (2014) 56:133–142

26. Kandanur, S.S., Rafiee, M.A., Yavari, F., Schrameyer, M., Yu, 29. Liu, W.W., Chai, S.P., Mohamed, A.R., Hashim, U.: Synthesis
Z.Z., Blanchet, T.A., Koratkar, N.: Suppression of wear in and characterization of graphene and carbon nanotubes: a review
graphene polymer composites. Carbon 50, 3178–3183 (2012) on the past and recent developments. J. Ind. Eng. Chem. 20,
27. Liu, J., Tang, J., Gooding, J.J.: Strategies for chemical modifi- 1171–1185 (2014)
cation of graphene and applications of chemically modified 30. Kalogeras, I.M., Hagg Lobland, H.E.: The nature of the glassy
graphene. J. Mater. Chem. 22, 12435–12452 (2012) state: structure and transitions. J. Mater. Educ. 34, 69–94 (2012)
28. Du, W.C., Jiang, X.Q., Zhu, L.H.: From graphite to graphene: 31. Salgado, J., Parajó, J.J., Fernandez, J., Villanueva, M.: Long-term
direct liquid-phase exfoliation of graphite to produce single- and thermal stability of some 1-butyl-1-methylpyrrolidinium ionic
few-layered pristine graphene. J. Mater. Chem. A 1, liquids. J. Chem. Thermodyn. 34, 69–94 (2014)
10592–10606 (2013)

123

You might also like